Open Access
How to translate text using browser tools
23 September 2009 Critical Function of AP-2gamma/TCFAP2C in Mouse Embryonic Germ Cell Maintenance
Susanne Weber, Dawid Eckert, Daniel Nettersheim, Ad J.M. Gillis, Sabine Schäfer, Peter Kuckenberg, Julia Ehlermann, Uwe Werling, Katharina Biermann, Leendert H.J. Looijenga, Hubert Schorle
Author Affiliations +
Abstract

Formation of the germ cell lineage involves multiple processes, including repression of somatic differentiation and reacquisition of pluripotency as well as a unique epigenetic constitution. The transcriptional regulator Prdm1 has been identified as a main coordinator of this process, controlling epigenetic modification and gene expression. Here we report on the expression pattern of the transcription factor Tcfap2c, a putative downstream target of Prdm1, during normal mouse embryogenesis and the consequences of its specific loss in primordial germ cells (PGCs) and their derivatives. Tcfap2c is expressed in PGCs from Embryonic Day 7.25 (E 7.25) up to E 12.5, and targeted disruption resulted in sterile animals, both male and female. In the mutant animals, PGCs were specified but were lost around E 8.0. PGCs generated in vitro from embryonic stem cells lacking TCFAP2C displayed induction of Prdm1 and Dppa3. Upregulation of Hoxa1, Hoxb1, and T together with lack of expression of germ cell markers such Nanos3, Dazl, and Mutyh suggested that the somatic gene program is induced in TCFAP2C-deficient PGCs. Repression of TCFAP2C in TCam-2, a human PGC-resembling seminoma cell line, resulted in specific upregulation of HOXA1, HOXB1, MYOD1, and HAND1, indicative of mesodermal differentiation. Expression of genes indicative of ectodermal, endodermal, or extraembryonic differentiation, as well as the finding of no change to epigenetic modifications, suggested control by other factors. Our results implicate Tcfap2c as an important effector of Prdm1 activity that is required for PGC maintenance, most likely mediating Prdm1-induced suppression of mesodermal differentiation.

INTRODUCTION

Primordial germ cells (PGCs) are the progenitors of gametes and are specified within the proximal epiblast around Embryonic Day 6.5 (E 6.5) in mice through signals from extraembryonic tissues [1, 2]. At E 7.25, PGCs can be identified due to their alkaline phosphatase reactivity as a cluster of approximately 45 cells at the base of the developing allantois [3, 4]. PGCs migrate into the embryo proper along the hindgut to colonize the developing genital ridges that will differentiate to become testes or ovaries [5, 6].

Compared to their neighboring somatic cells, PGCs are characterized by high expression of Ifitm3 and Dppa3 (previously known as “fragilis” and “stella,” respectively). In contrast, genes indicative of somatic differentiation, such as Hoxb1 and Hoxa1, are repressed. It has been proposed that repression of somatic differentiation is one mechanism by which the PGCs escape from the somatic fate and retain pluripotency. Recently, the transcriptional repressor BLIMP1 (B-lymphocyte-induced maturation protein-1, official symbol PRDM1) [7] has been identified as a key factor controlling the specification of PGCs [8]. Upon loss of Prdm1, PGC specification is blocked at an early stage, resulting in lack of activation of PGC-specific genes and derepression of HOX genes leading to induction of somatic differentiation [8]. Global gene expression profiling of the initial 48 h of germ cell development was used to decipher the regulatory network governed by PRDM1 to specify PGCs [9]. Genes found highly upregulated in early PGCs following PRDM1 induction comprise several transcriptional regulators, among them Tcfap2c, Prdm14, Sox3, Sox17, Elf3, Elk1, Isl2, Mycn, Klf2, 4932441K18Rik (Fiat), Sp8, and Smad3 [9].

The precise roles these PRDM1 targets play in the suppression of somatic differentiation have so far remained unknown. Here, we focused on one transcriptional regulator identified in this study, Tcfap2c (also termed AP-2γ), and analyzed the consequences of a loss of function mutation in vivo and in vitro.

Transcription factor Tcfap2c belongs to a family of five closely related genes that are expressed during embryogenesis [10, 11]. Disruption of Tcfap2c leads to midgestation lethality due to a defect in the trophectoderm cells [12, 13]. Within the embryo proper, Tcfap2c is expressed in a variety of tissues, including PGCs [14]. In humans, Tcfap2c is expressed in fetal germ cells from Week 12 through Week 37 of pregnancy, while it is downregulated and subsequently absent in healthy adult testes [15, 16]. In spite of this highly regulated temporal and spatial expression pattern, nothing is known about the function of Tcfap2c in germ cell development. Therefore, using both in vitro and in vivo models, we have investigated the role of TCFAP2C during PGC development.

MATERIALS AND METHODS

Genotyping

DNA was prepared from tail biopsies or embryonic tissues using standard procedures. Genotyping of Tcfap2c alleles was done as described previously [12]. Primers used to detect the cre alleles were 5′-CCA CGA CCA AGT GAC AGC AAT G-3′ and 5′-CAG AGA CGG AAA TCC ATC GCT C-3′, both of which amplify a 373bp DNA fragment. Cycle conditions were 94°C for 45 sec, 56°C for 30 sec, and 72°C for 45 sec for 35 cycles.

Whole Mount Alkaline Phosphatase Staining of Embryos

Embryos were dissected in PBS, fixed in 4% (w/v) paraformaldehyde (PFA) in PBS for 10 min at room temperature, washed twice in PBS, and then incubated in freshly prepared alkaline phosphatase (AP)-staining solution, i.e., 250 mM Tris-maleic acid, pH 9, 0.4 mg/ml α-naphtylphosphate (Sigma-Aldrich, Taufkirchen, Germany), and 1 mg/ml Fast Red (Sigma-Aldrich), for 30 min in the dark. Embryos were cleared in 40% glycerol for 1 h and 80% glycerol overnight at 4°C. Embryos were photographed using a binocular (Leica, Bensheim, Germany) and a Zeiss Axiocam digital camera system (Carl Zeiss, Jena, Germany). Files were processed using Adobe Photoshop Elements and Illustrator software (Adobe, San Jose, CA).

Animal Care and Husbandry

All animals were housed at 25°C with a 12L:12D photoperiod and were given water and a standard rodent diet ad libitum. All animal experiments were approved by the Landesamt fuer Natur, Umwelt und Verbraucherschutz Nordrhein-Westfalen (#8.87–50.10.31.08). The experiments were conducted in accordance with the International Guiding Principles for Biomedical Research Involving Animals as promulgated by the Society for the Study of Reproduction. To obtain Tcfap2c-deficient mice, we carried out a multistep mating scheme where mice heterozygous for the Tcfap2c allele (Tcfap2c+/−) were crossed with cre-expressing mice (cre+). The resulting Tcfap2c heterozygous cre double transgenic animals (Tcfap2c+/−; cre+) were crossed with Tcfap2c flox/flox mice to receive the null mutants. The day on which a copulation plug was found was defined as E 0.5.

Laser Microdissection

P.A.L.M. membrane slides (Carl-Zeiss MicroImaging, Göttingen, Germany) were irradiated with ultraviolet light for 30 min. Paraffin-embedded tissue was sectioned to 10 μm thickness and stained with haematoxylin and eosin using standard procedures. Seminiferous tubules were dissected using P.A.L.M. Microlaser Technology (Carl-Zeiss MicroImaging). Dissected tissue was collected in DNA lysis buffer (1 mM EDTA pH 8, 10 mM Tris-HCl pH 8, 0.5% Tween20, and 100 μg/ml proteinase K) and incubated at 37°C for 24 h, heated to 95°C for 12 min the next day, and used as a template for PCR.

RT-PCR

Total RNA was extracted using the Qiagen RNeasy Kit (Qiagen, Hilden, Germany); 100 ng was used for cDNA synthesis (SuperScriptIII; Invitrogen, Karlsruhe, Germany). PCR was performed with the primers listed in Supplemental Table S1 (all the Supplemental Data are available online at www.biolreprod.org). Thirty-five cycles (30 in the case of β-actin) were run using a standard protocol.

Quantitative RT-PCR

High-quality total RNA was extracted from the cell lines using an RNAqueous-4PCR kit (Ambion/Applied Biosystems, Nieuwerkerk a/d IJssel, the Netherlands) according to the manufacturer's instructions. Samples were treated with DNase and checked for residual DNA contamination by PCR using a primer set for SOX2, which is a single exon gene (see Supplemental Table S1). Complementary DNA was generated as described previously [17]. Quantitative PCR was performed using the 7900HT Fast Real-Time PCR System (Applied Biosystems, Foster City, CA). For the detection of expression of a set of genes related to various differentiation lineages, a number of primer sets were used. Primer sequences are given in Supplemental Table S1. Target mRNA was quantified relative to HPRT transcripts as follows: target mRNA value = 2(mean Ct HPRT − Ct target) [17].

Whole Mount In Situ Hybridization

Whole mount in situ hybridizations were performed as described previously [18]. Tcfap2c-full length cDNA subcloned in pBS vector was used for generation of the probe. Embryos were photographed using a binocular (Leica) and a Zeiss Axiocam digital camera system (Carl Zeiss). Files were processed using Adobe Photoshop Elements and Illustrator software (Adobe).

Histology and Imunohistochemistry

Embryos or organs were fixed in 4% neutral buffered formalin at 4°C overnight and embedded in paraffin. Sections of 1–3 μm thickness were stained with hematoxilin and eosin using standard procedures. For immunohistochemistry, sections were incubated using the following primary antibodies: YBX2 (MSY2) (1:2000, kindly provided by Dr. R.M. Schultz, Department of Biology, University of Pennsylvania, Philadelphia, PA), GATA1 (1:200, Santa Cruz, Heidelberg, Germany), GCNA (kindly provided by Dr. G. Enders, University of Kansas, Kansas City, KS), NOS1 (nNOS) (1:200, BD Biosciences, Heidelberg, Germany), TCFAP2C (Clone 6E4, 1:250, kindly provided by Dr. H. Hurst, Cancer Research UK, London, UK), for 1 h at 37°C except YBX2 (4°C, overnight). The secondary antibodies were anti-rabbit (1:500; DAKO, Hamburg, Germany) and anti-rat (1:200; DAKO) for 30 min at room temperature. Signals were visualized using Vectastain ABC Kit (Vector Laboratories, Axxora Deutchland, Lörrach, Germany). Sections were photographed using Diskus software (Diskus, Hilden, Germany). Files were processed using Adobe Photoshop Elements 5 and Illustrator software (Adobe).

Isolation and Immunostaining of Primordial Germ Cells

Embryos at E 7.5 were collected in Dulbecco modified Eagle medium (DMEM) (Gibco, Karlsruhe, Germany) supplemented with 0.5% BSA. Embryonic fragments from the base of the allantois were isolated using a microscalpel and incubated in 0.05% trypsin and 0.5 mM EDTA for 7 min, followed by dissociation into single cells by trituration with a mouth pipette. Dissociated single cells were pipetted onto poly-l-lysine-coated glass slides and fixed in 2% PFA. Cells were permeabilized in 0.1% Triton-X-100/PBS, and primary antibody incubation was carried out overnight at 4°C. The following antibodies and dilutions were used: anti-Tcfap2c (H-77, 1:100; Santa Cruz Biotechnology, Inc.), anti-Blimp-1 (Prdm-1 3H2-E8, 1:20; Affinity Bioreagents, Golden, CO), Alexa-488 (1:500; Invitrogen, Karlsruhe, Germany), and Alexa-594 (1:500; Invitrogen).

Detection of Cell Death

LysoTracker (Invitrogen, Germany) staining was performed as previously described [19]. Acidic compartments within apoptotic cells and regular cells engulfing debris from apoptotic cells are labeled. For these analyses, E 8.0 mouse embryos were prepared, incubated in LysoTracker solution, i.e., 5 μl LysoTracker, 1 ml DMEM (without phenol-red), and 2 mg/ml BSA (Sigma Aldrich, Munich, Germany), at 37°C for 40 min, washed twice in DMEM/BSA, twice in PBS, and fixed with 4% PFA in PBS for 2 h at 4°C. Whole mounts were incubated with 4′,6-diamidino-2-phenylindole (DAPI) (Sigma Aldrich), and the staining was analyzed using a Zeiss-ApoTome (Carl Zeiss) with a fluorescence module and appropriate filter set.

Differentiation and Selection of PGC from Embryoid Body Cultures

Essentially, the procedure was performed as described previously [20]. For embryoid body (EB) formation, ES cells were washed twice in calcium- and magnesium-free Dulbecco phosphate solution (DPBS, Invitrogen, Germany) and dissociated with 0.05% trypsin/EDTA solution (PAN-Biotech, Aidenbach, Germany). Trypsin was neutralized with DMEM (Invitrogen) containing 15% fetal bovine serum (Hyclone, Bonn, Germany). To remove feeder cells, the ES cells were plated twice on a 10-cm culture dish (TPP, Trasadingen, Switzerland) with 0.2% gelatin. After 45 min, nonadherent ES cells were collected, and hanging drop culture (750 ES cells per 25 μL) on lids of dishes filled with PBS was initiated. After 3 days in hanging drop culture, the resulting EBs were transferred to bacterial dishes (10–15 EB per dish) to prevent attachment. Every second day, one-third of the medium was replaced. SSEA-1 positive cell fractions were obtained using MACS-Beads and MACS-LS columns (Miltenyi Biotec, Cologne, Germany) according to the manufacturer's protocol. Five microliters anti-SSEA-1 antibody (0.1 mg/ml; R&D Systems, Heidelberg, Germany) per 2 × 105 total cells were used.

RESULTS

Tcfap2c Is Expressed in Primordial Germ Cells and Gonocytes

First the expression of Tcfap2c in PGCs and gonocytes during embryogenesis was analyzed. Using immunohistochemistry, we detected Tcfap2c-positive cells starting from the early bud stage at E 7.25 (the earliest stage analyzed) at the base of the allantois, the region where PGCs are located at this developmental stage (Fig. 1, A and B, and Supplemental Fig. S1A). To confirm that the cells expressing Tcfap2c were germ cells, the base of the allantois (box in Fig. 1A) was dissected, trypsinized, and labeled with antibodies detecting the PGC markers Prdm1 and Tcfap2c (Fig. 1, C1 and C2). Prdm1 and Tcfap2c were coexpressed in these cells (Fig.1, C3 and C4), suggesting that Tcfap2c is expressed in PGCs. Next, we extracted RNA from the dissected allantoises at E 7.25. RT-PCR analyses showed that the markers for PGCs Kit and Dnd1 (previous symbol Ter) were coexpressed with Tcfap2c (Supplemental Fig. S1B). These results are in agreement with data obtained from transgenic mice expressing PRDM1-mVenus, where transgene expression was detected in Tcfap2c-positive cells at the base of the allantois in E 7.25 embryos [9]. Taken together, these data demonstrate that Tcfap2c is expressed in PGCs shortly after specification. At E 8.25, PGCs have begun to migrate into the embryo along the hindgut and express the marker Dppa3. At this stage, Tcfap2c expression (Fig. 1D) colocalizes with the signal obtained for Dppa3 (Fig. 1, E1–E3). Using in situ hybridization, we could detect Tcfap2c transcripts in the region of the hindgut at E 9.5. The observed pepperlike pattern is typical for migrating PGCs (Fig. 1F). At E 10.5 and E 11.5, we detected Tcfap2c transcripts in postmigratory PGCs in the genital ridges (Fig. 1, G and H). Staining of paraffin sections with a TCFAP2C-specific antibody confirmed this pattern from E 11.5 to E 12.5 (Fig. 1, I and K). TCFAP2C protein or RNA could neither be detected in germ cells in genital ridges after E 12.5 nor in adult testes or ovaries (Fig. 1L and Supplemental Fig. S1, C–E). These results are in agreement with data obtained from humans where expression of Tcfap2c was observed in primordial germ cells and gonocytes, but not in prespermatogonia and spermatogonia [15, 16]. These results prompted us to investigate the role of Tcfap2c in germ cell development.

FIG. 1.

Tcfap2c is expressed in primordial germ cells and gonocytes. Immunoflourescent detection of TCFAP2C at (AC) E 7.25 and (D and E) E 8.25. Boxed area in A is enlarged in B. C1C4) Double labeling of cell suspension generated by dissecting boxed area from A using (C1) PRDM1 and (C2) TCFAP2C antibodies; (C3) counterstaining with DAPI; and (C4) merge of C1C3. D) Detection of TCFAP2C expression at E 8.25 along the forming hindgut. Boxed area in D is enlarged in E1 and E2 and shows the fluorescent signal for DPPA3 and TCFAP2C, respectively. E3) Merge of E1 and E2, demonstrating the coexpression of both markers. Messenger RNA in situ hybridization of Tcfap2c at (F) E 9.5, (G) E 10.5, and (H) E 11.5. Arrows in B, C1, C2, E3, F, and G indicate primordial germ cells. IL) Antibody detecting TCFAP2C in sections at (I) E 11.5, (K) E 12.5, and (L) E 15.5. a, allantois; gr, genital ridge; lb, limb bud. Bar = 50 μm.

i0006-3363-82-1-214-f01.tif

Tcfap2c Deficient Mice Fail to Reproduce

A complete loss of Tcfap2c is lethal to the embryo around E 6.5 due to a placenta defect and thus precludes the analysis of PGC development [12, 13]. To overcome this difficulty, we took advantage of transgenic mice bearing conditionally mutated Tfap2c alleles [21]. These Tcfap2c flox/flox mice [21] were bred to either Mox2-cre mice [22] expressing the Cre recombinase from E 6 on in the embryo proper or to Alpl-cre mice [23] expressing the Cre protein in PGCs. Both, Tcfap2c flox/flox:Mox2-cre and Tcfap2c flox/flox:Alpl-cre animals developed normally but failed to reproduce and displayed a similar phenotype with respect to germ cell development. Testes and ovaries of the mutant animals were dramatically reduced in size compared to wild-type controls (Fig. 2, A and B). In female animals, lack of germ cells in ovaries was demonstrated using YBX2 as marker for oocytes (Fig. 2, D vs. C). In male animals, atrophy of the seminiferous tubules was apparent (Fig. 2, F vs. E), and immunohistochemistry demonstrated that mutant testes did not contain any germ cells as judged by the marker YBX2 (Fig. 2, H vs. G) [24]. In contrast, GATA1-positive Sertoli cells were present (Fig. 2, K vs. I) [25] as well as Leydig cells that could be visualized with anti-NOS1 (Fig. 2, M vs. L) [26]. Of note, Leydig cell hyperplasia (Fig. 2, lch) was obvious in the mutant testis, which most likely is a secondary effect due to lack of germ cells. In some animals, the germ cell atrophy was incomplete. PCR analysis of laser microdissected tissue revealed that the Cre-mediated loss of Tcfap2c had not occurred in such testicular areas (Supplemental Fig. S2), suggesting that the incomplete atrophy was due to an incomplete Cre-mediated excision. Indeed, the recombination frequency of the Alpl-cre allele has been reported to reach 72% [27]. These results demonstrate that loss of Tcfap2c leads to sterility.

FIG. 2.

Loss of Tcfap2c leads to lack of germ cells. A) Testes and (B) ovaries of control (Wt) and Tcfap2cflox/flox:Mox2-cre (mut) mice. Bar = 2 mm. MSY-2 immunohistochemistry on ovarian tissue of (C) control and (D) Tcfap2cflox/flox:Mox2-cre. Sections of (C, E, G, I, L) control and (D, F, H, K, M) Tcfap2cflox/flox:Alpl-cre mutants stained with hematoxylin and eosin (H&E) or the antibody indicated. Specimen and sections were prepared from 3-mo-old animals. st, seminiferous tubules; lch, Leydig cell hyperplasia. Bar = 200 μm (EM) and 500 μm (C and D).

i0006-3363-82-1-214-f02.tif

Tcfap2c Deficient Germ Cells Are Lost Shortly after Specification

Because the temporal expression of Tcfap2c in the germ cell lineage is restricted to PGCs and gonocytes, we tried to determine at which stage during early germ cell development might Tcfap2c be essential. To address this question, we crossed the Tcfap2c flox/flox mice with Sox2-cre transgenic mice. The Sox2-cre transgene is expressed from E 5.0 in all cells of the epiblast [28]. Therefore, Sox2-cre transgenic mice could be used in a manner similar to the Mox2-cre-transgenic mice to overcome the placenta defect caused by complete loss of Tcfap2c. However, Cre-mediated recombination is more efficient compared to Mox2-cre or Alpl-cre, thus, precluding a mosaic recombination within a tissue (unpublished data and [28]).

Germ cell development of Tfap2c flox/flox:Sox2-cre or control embryos was monitored using AP staining. PGCs could be detected at the primitive streak stage (E 7.25) at the base of the allantois both in the mutants and in the littermate controls (Fig. 3, A and B, arrows). The total number appeared not to be altered. However, at E 8.0, the numbers of PGCs in mutant mice were drastically reduced compared to control mice. In mutant animals, only a few PGCs were detectable near the base of the allantois, and none had initiated migration in contrast to the wild types (Fig. 3, D vs. C, arrows). At E 9.5, no migrating PGCs could be detected along the hindgut in mutant animals in contrast to controls (Fig. 3, E vs. F, arrows). Moreover, at E 10.5 the genital ridges of mutant animals lacked AP staining indicative of PGCs (Fig. 3, H vs. G, arrows,). At E 12.5, wild-type genital ridges contained large amounts of germ cells as indicated by the intense staining. In contrast, the genital ridges of Tcfap2c flox/flox: Sox2-cre mice were completely devoid of the AP signal (Fig. 3, J vs. I). Immunohistochemical staining using the germ cell marker GCNA at E 11.5 confirmed that mutant animals lacked germ cells (Fig. 3, K and L). These results demonstrate that the loss of Tcfap2cin PGCs interfered with early germ cell development from E 8.0 onward.

FIG. 3.

Tcfap2c is required for germ cells. Alkaline phosphatase (AP) whole mount staining of control (wt) and mutant (Tcfap2cflox/flox:Sox2-cre, mut) (A and B) whole embryos, (C and H) trunk, and (I and J) genital ridge at the embryonic age indicated. The arrows indicate the position of AP-positive germ cells. K, L) Antibody to GCNA detection of germ cells in sections of genital ridges of E 11.5 control (WT) and mutant (mut). gr, genital ridge; kd, kidney; mn, mesonephros. Bar =100 μm (CJ) and 25 μm (K and L).

i0006-3363-82-1-214-f03.tif

No Enhanced Apoptosis in Tcfap2c-Deficient Germ Cells

To investigate a possible loss of PGCs due to apoptosis, as observed in Pou5f1-deficient mice [29], LysoTracker analysis [19] of E 8.0 Tcfap2c flox/flox:Sox2-cre embryos was performed. LysoTracker signal in the region between the allantois and the developing hindgut, where PGCs are localized at this point of embryonic development, was not enhanced in the mutants (Fig. 4A) compared to wild types (Fig. 4B). Apoptosis in E 8.0 head-fold mesenchyme (Fig. 4C) was easily detected and served as positive control. The bright-field images were merged with the signal obtained by the LysoTracker to show the embryo and allantois (Fig. 4, A′–C′). Statistical analysis of this experiment showed that the differences between the numbers of LysoTracker positive cells in wild type (11 ± 2.55) compared to Tcfap2c-deficient (11.74 ± 2.73) embryos were not significantlt different ( Fig. 4D; P = 0.722, Student t-test). In addition, no difference in apoptosis could be detected using whole mount TUNEL-assays (data not shown) [30]. Therefore, cell loss by apoptosis is unlikely to be the mechanism responsible for the lack of germ cells in Tcfap2c mutant animals.

FIG. 4.

Cell death is not enhanced in mutant PGCs. Photomicrography of LysoTracker analysis (red signal, LT) of (A and A′) mutant (mut) and (B and B′) wild type (Wt) animals at E 8.0. The allantois (a) and posterior end of embryo is shown. C, C′) Apoptosis in the head mesenchyme of wild-type animal serves as positive control. AC) Counterstaining with DAPI. A′–C′) Merge of bright field (BF) and lysotracker (red, LT) signal. Bar = 100 μm. D) Graph showing the number and standard deviation of cells positive in the posterior end of embryos (n = 8) in the LysoTracker study for wild type (WT; 11 ± 2.55) and Tcfap2c mutants (11.74 ± 2.73).

i0006-3363-82-1-214-f04.tif

PGCs Derived from Tcfap2c-Deficient Embryoid Body Cultures Upregulate Somatic Markers

The defects observed in PGCs of Tcfap2c mutants are a phenocopy of the defects observed in Prdm1-deficient mice. Loss of Prdm1 leads to induction of somatic differentiation as demonstrated by the upregulation of Hoxb1 [8]. We used an in vitro model for PGC differentiation to test whether lack of Tcfap2c would likewise result in Hoxb1 gene activation. PGCs are known to be present in EB cultures derived from murine embryonic stem cells. Stage-specific embryonic antigen (SSEA1)-positive cells obtained from EB cultures represent candidate PGCs [20]. Tcfap2c-deficient ES cells were generated from Tcfap2c-deficient blastocysts derived from a Tcfap2c +/− line that was bred into the 129Sv background (our unpublished results). EB cultures were initiated from wild-type ES cells as well as from Tcfap2c-deficient ES cells. We used two different lines of Tcfap2c-deficient ES cells, yielding similar results. After 3, 5, and 7 days, the EBs were trypsinized, and SSEA1-positive cells (the nascent PGCs) were isolated using immunomagnetic bead sorting. As published [20], expression of Pou5f1 remained constant in SSEA1-positive cells while Pou5f1 levels declined in the SSEA1-negative population (data not shown). The markers for early PGCs, Prdm1 and Dppa3, could be detected in both wild-type and Tcfap2c-deficient cultures (Fig. 5A, top), indicating that PGCs had been specified in both genotypes. However, the markers for mature PGCs (Nanos3, Dazl, and Mutvh), which are also expressed in ES cells, decreased significantly in SSEA1-positive cells of Tcfap2c-deficient EB cultures compared to wild-type controls over the time observed (Fig. 5A, middle). Notably, the markers for somatic differentiation could either be detected only in cells from Tcfap2c-deficient EB cultures (Hoxb1) or were induced stronger and earlier in this assay (Hoxa1, T, and Snai1) (Fig. 5A, lower). Taken together, these data suggest, that in Tcfap2c-deficient EB cultures, PGCs are specified but further germ cell-specific differentiation is impaired and somatic differentiation is initiated.

FIG. 5.

Lack of Tcfap2c results in upregulation of somatic markers. A) Wild-type (Wt) and Tcfap2c-deficient (Tcfap2c −/−) ES cells were subjected to hanging drop cultures to induce embryoid body formation (EB) for 3 (d3), 5 (d5), and 7 (d7) days. SSEA1-positive cells were isolated after trypsinization, and RT-PCR analyses of expression of the genes indicated was performed. Primer sequences are given in Supplemental Table S1. The experiments were repeated 10 times. B) TCam-2 cells were treated with siRNA mixture to repress TCFAP2C (kd) and analyzed with respect to untreated cells (u) or cells treated with unspecific siRNA (scr). RT-PCR was used to detect the expression of the genes indicated; (−) indicates negative control without RT. Primer sequences are given in Supplemental Table S1. The experiments were repeated three times.

i0006-3363-82-1-214-f05.tif

Loss of TCFAP2C Leads to Upregulation of Mesoderm Markers

Kurimoto et al. [9] demonstrated that Prdm1 regulates germ cell specification by the inhibition of somatic differentiation, specifically by repressing HOX cluster activation, epithelial-mesenchymal transition, cell cycle progression, and the DNA-methyltransferase machinery. Therefore, we analyzed the role of Tcfap2c in the genetic program orchestrated by Prdm1. As an in vitro model, we used TCam-2 [31], a cell line derived from a human seminoma that displays significant characteristics of PGC. Indeed, TCam-2 cells express several PGC markers such as PRDM1 and PRMT5 [32], and TCFAP2C, SOX17, KIT, DPPA3, NANOG, and POU5F1 [3335]. In addition, TCam-2 cells exhibit symmetric dimethylation of arginine-3 on histones H2A and/or H4 tails (H2A/H4R3me2s) [32]. This modification has been demonstrated to be specific for PGCs and is directed by the interaction of Prdm1 and the arginine-specific histone methyltransferase Prmt5 [36].

First, we tested whether TCam-2 cells could serve as an in vitro model for PRDM1-mediated repression of the somatic program on the genetic and epigenetic level. Here, short interfering RNA (siRNA)-mediated knockdown of PRDM1 led to a reduction of TCFAP2C expression levels and a concomitant upregulation of HOXB1. In addition, expression of the PRDM1 target gene MYC was induced while the expression of NANOG remained unaffected (Supplemental Fig. S3A). Western blot analyses demonstrated that the symmetric dimethylation on histone H2A/H4 was lost after PRDM1 knockdown (Supplemental Fig. S3B). These data indicated that TCam-2 seminoma cells can serve as a model to study PRDM1-mediated gene programs in PGCs and further corroborated that TCFAP2C represents a PRDM1 target gene.

Next, we established a siRNA-mediated knockdown of TCFAP2C in TCam-2 cells to analyze TCFAP2C-dependent genetic programs in more detail. We compared the knockdown efficiency of three single siRNAs to their mixture (as suggested by the supplier) and detected the strongest inhibition of TCFAP2C RNA and protein when using the mixture (Supplemental Figure S3C). Next, we performed quantitative RT analyses of 31 markers (see Supplemental Table S1). As depicted in Supplemental Fig. 3D, expression of pluripotency genes (NANOG, POU5F1, SOX17, SOX2, UTF1, and REXO1), genes involved in DNA de novo methylation (DNMT3A, DNMT3B, and DNMT3L), signaling (SMAD4, SMAD7, and STAT5A), and various markers for ectoderm and endoderm differentiation remained unaffected after downregulation of TCFAP2C. However, the genes indicative of mesoderm differentiation, MYOD1 and HAND1, showed strong upregulation. Also GATA2, a marker found in mesoderm-derived hematopoietic progenitors [37], was upregulated. Using conventional RT-PCR, induction of MYOD1 and HAND1 as well as HOXA1 and HOXB1 could be detected in TCam-2. Induction of T was only weak compared to the results of the EB cultures while the levels of Snai1 were not altered in this context. Of note, levels of PRDM1 remained unaffected by TCFAP2C knockdown (Fig. 5B). Taken together, we demonstrated that repression of TCFAP2C leads to upregulation of HOXB1 and the mesoderm markers MYOD1 and HAND1 while markers of other germ layers, cell cycle progression, signaling, and DNA methyltransferases remain unaffected. These results therefore suggest that TCFAP2C is involved in the repression of a set of genes inducing mesodermal differentiation.

In summary we propose a model, wherein Tcfap2c is an effector of PRDM1-mediated repression of somatic differentiation, in particular differentiation into mesoderm. Pluripotency-associated markers and genes involved in epigenetic reprogramming are not affected.

DISCUSSION

In this study, we show that Tcfap2c is specifically expressed in PGCs and gonocytes from E 7.25 to E 12.5 of murine development, suggesting a role in germ cell biology. Using a conditional deletion approach for Tcfap2c, we show that PGCs are lost shortly after specification that results in sterile animals, independent of sex. In addition, we show that repression of Tcfap2c leads to upregulation of Hoxb1, Hand1, and Myod expression, which is indicative of mesoderm differentiation. The expression level of the master coordinator of germ cell development, Prdm1, remains unaffected, supporting the idea that Tcfap2c acts as a downstream mediator in this pathway.

In PGCs, Tcfap2c expression is induced shortly after their specification. Secretory bone morphogenetic proteins (Bmp) Bmp 4, Bmp8b, and Bmp2 and the downstream signaling molecules of the SMAD pathway are involved in the specification of PGCs [38]. The exact molecular mechanism, however, has not been elucidated. Interestingly, in Xenopus epidermis and rat neural crest, Bmp signaling has been reported to induce Tcfap2c expression, suggesting a Bmp-mediated induction of Tcfap2c in nascent PGCs as a possible mechanism [3942].

Does the lack of TCFAP2C interfere with PGC specification? Experimental data argue against this scenario. If Tcfap2c were required to mediate Bmp and Smad signaling, PGC specification would not take place. However, AP-positive PGCs were detected at E 7.0 in Tcfap2c-mutant mice. Also, SSEA1-positive cells from Tcfap2c-deficient EB cultures showed induction of Prdm1 and Dppa3, early markers for PGCs. In addition, expression levels of SMAD4 and SMAD7, the downstream effectors of BMP signaling, remained unaffected in TCam-2 cells after siRNA mediated knockdown of TCFAP2C. However, at E 8.0 PGCs in the mutants appeared reduced in number and had not initiated migration into the embryo suggesting an impaired proliferation of the founding population and/or an altered migratory behavior.

Loss of germ cells in the Tcfap2c mutants may be accomplished by two mechanisms: cell death by apoptosis or differentiation into somatic cells. PGCs harboring mutations in Pou5f1 or Kit undergo apoptosis; hence, these factors are regarded as survival factors for germ cells during migration and proliferation [29, 43]. We showed that levels of Pou5f1 and Kit were not affected by lack of Tcfap2c both in EB differentiation assays and in Tcfap2c knockdown experiments. In addition, the number of apoptotic cells is not enhanced in Tcfap2c-deficient embryos. Therefore, we conclude that Tcfap2c and Pou5f1 or Kit act in different pathways during PGC development and migration.

However, parallels to the phenotype caused by loss of Prdm1 are apparent [8]. Prdm1-deficient PGCs are specified but fail to migrate, and they are prone to uncontrolled somatic differentiation [8]. Indeed, in PGCs, Tcfap2c expression has been shown to depend on and correlate with Prdm1 expression [9]. In PGCs, Prdm1 is required for the repression of developmental programs such as the epithelial-mesenchymal transition, repression of HOX cluster activation, repression of cell cycle progression, and the repression of the DNA-methyltransferase machinery [9]. Using an EB-based differentiation protocol, we demonstrate induction of Prdm1 and Dppa3 in candidate PGCs. In the literature, established PGCs have been characterized as Dppa3-positive and Hoxb1-negative cells [44, 45]. Lack of Tcfap2c in EB cultures, however, leads to induction of Hoxb1 in DPPA3 (and PRDM1)-positive cells. It has been shown, that embryos of the late primitive streak stage PGCs coexpress Prdm1 and Hoxb1 during specification [9]. Genetic lineage-tracing experiments revealed that all Prdm1-expressing cells at early stages contribute almost invariably to DPPA3-positive PGCs, which eventually repress Hoxb1 [8]. The fact that the SSEA1-positive cells of the EB cultures coexpress Dppa3 and Prdm1 indicates that these PGCs have been correctly specified. Reexpression of Hoxb1 after 3 to 5 days in the Tcfap2c-deficient EB cultures even in the presence of PRDM1 protein suggests that Tcfap2c acts downstream of Prdm1, and Prdm1-induced repression of Hoxb1 might be mediated by Tcfap2c. Results from siRNA-mediated knockdown in TCam-2 cells further corroborate these conclusions. Here, TCFAP2C knockdown lead to reexpression of Hoxb1 in the presence of PRDM1 protein as well. In particular, Tcfap2c seems to specifically suppress mesoderm differentiation, as demonstrated by upregulation of HAND1 and MYOD following TCFAP2C downregulation. In contrast, the expression levels of markers for ectoderm, endoderm, or extraembryonic differentiation were not affected by TCFAP2C, suggesting that these differentiation pathways might be controlled by other downstream regulators of PRDM1. Also the repression of DNA methyltransferases DNMT3A, B, and L in PGCs, which results in a genome-wide DNA demethylation [46], is not affected by downregulation of TCFAP2C. Recent results suggest that Prdm14 controlled epigenetic reprogramming and reacquisition of pluripotency in PGCs represents a Prdm1-independent pathway [47]. Based on our data, we propose that Tcfap2c is acting downstream of Prdm1 and represses one aspect of somatic (i.e., mesoderm) differentiation in PGCs.

The expression pattern described for Tcfap2c in mice is in accordance with the expression of TCFAP2C in human fetal germ cells [15, 16], suggesting that this protein has conserved functions in both species. Moreover, TCFAP2C has been shown to be expressed in undifferentiated germ cell tumors, namely in IGCNU/CIS (intratubular germ cell neoplasia unclassified, carcinoma in situ) and seminomas [15, 16]. The absence of TCFAP2C in differentiated germ cell tumors (e.g., embryonal carcinoma and teratocarcinoma) provokes speculation as to whether the loss of TCFAP2C expression in germ cell tumors may enable the activation of a somatic developmental program, leading to differentiation into embryonal carcinoma and teratocarcinoma.

Acknowledgements

We thank R. Jäger for critically reading the manuscript; I. Heim, A. Jäger, G. Beine, A. Egert, S. Steiner, and B. Reddemann for excellent technical assistance; Dr. G. Enders for the GCNA antibody; Dr. R. Schultz for the MSY2 antibody; and Dr. H. Hurst for the Tcfap2c antibody.

REFERENCES

1.

KA Lawson, NR Dunn, BA Roelen, LM Zeinstra, AM Davis, CV Wright, JP Korving, and BL Hogan . Bmp4 is required for the generation of primordial germ cells in the mouse embryo. Genes Dev 1999. 13:424–436. Google Scholar

2.

KA Lawson and WJ Hage . Clonal analysis of the origin of primordial germ cells in the mouse. Ciba Found Symp 1994. 182:68–84;. discussion. 84–91. Google Scholar

3.

AD Chiquoine . The identification, origin, and migration of the primordial germ cells in the mouse embryo. Anat Rec 1954. 118:135–146. Google Scholar

4.

M Ginsburg, MH Snow, and A McLaren . Primordial germ cells in the mouse embryo during gastrulation. Development 1990. 110:521–528. Google Scholar

5.

R Anderson, TK Copeland, H Scholer, J Heasman, and C Wylie . The onset of germ cell migration in the mouse embryo. Mech Dev 2000. 91:61–68. Google Scholar

6.

KA Molyneaux, J Stallock, K Schaible, and C Wylie . Time-lapse analysis of living mouse germ cell migration. Dev Biol 2001. 240:488–498. Google Scholar

7.

CA Turner Jr, DH Mack, and MM Davis . Blimp-1, a novel zinc finger-containing protein that can drive the maturation of B lymphocytes into immunoglobulin-secreting cells. Cell 1994. 77:297–306. Google Scholar

8.

Y Ohinata, B Payer, D O'Carroll, K Ancelin, Y Ono, M Sano, SC Barton, T Obukhanych, M Nussenzweig, A Tarakhovsky, M Saitou, and MA Surani . Blimp1 is a critical determinant of the germ cell lineage in mice. Nature 2005. 436:207–213. Google Scholar

9.

K Kurimoto, Y Yabuta, Y Ohinata, M Shigeta, K Yamanaka, and M Saitou . Complex genome-wide transcription dynamics orchestrated by Blimp1 for the specification of the germ cell lineage in mice. Genes Dev 2008. 22:1617–1635. Google Scholar

10.

K Hilger-Eversheim, M Moser, H Schorle, and R Buettner . Regulatory roles of AP-2 transcription factors in vertebrate development, apoptosis and cell-cycle control. Gene 2000. 260:1–12. Google Scholar

11.

D Eckert, S Buhl, S Weber, R Jager, and H Schorle . The AP-2 family of transcription factors. Genome Biol 2005. 6:246..  Google Scholar

12.

U Werling and H Schorle . Transcription factor gene AP-2 gamma essential for early murine development. Mol Cell Biol 2002. 22:3149–3156. Google Scholar

13.

HJ Auman, T Nottoli, O Lakiza, Q Winger, S Donaldson, and T Williams . Transcription factor AP-2gamma is essential in the extra-embryonic lineages for early postimplantation development. Development 2002. 129:2733–2747. Google Scholar

14.

C Chazaud, M Oulad-Abdelghani, P Bouillet, D Decimo, P Chambon, and P Dolle . AP-2.2, a novel gene related to AP-2, is expressed in the forebrain, limbs and face during mouse embryogenesis. Mech Dev 1996. 54:83–94. Google Scholar

15.

K Pauls, R Jager, S Weber, E Wardelmann, A Koch, R Buttner, and H Schorle . Transcription factor AP-2gamma, a novel marker of gonocytes and seminomatous germ cell tumors. Int J Cancer 2005. 115:470–477. Google Scholar

16.

CE Hoei-Hansen, JE Nielsen, K Almstrup, SB Sonne, N Graem, NE Skakkebaek, H Leffers, and E Rajpert-De Meyts . Transcription factor AP-2gamma is a developmentally regulated marker of testicular carcinoma in situ and germ cell tumors. Clin Cancer Res 2004. 10:8521–8530. Google Scholar

17.

LH Looijenga, R Hersmus, AJ Gillis, R Pfundt, HJ Stoop, RJ van Gurp, J Veltman, HB Beverloo, E van Drunen, AG van Kessel, RR Pera, DT Schneider, B Summersgill, J Shipley, A McIntyre, P van der Spek, E Schoenmakers, and JW Oosterhuis . Genomic and expression profiling of human spermatocytic seminomas: primary spermatocyte as tumorigenic precursor and DMRT1 as candidate chromosome 9 gene. Cancer Res 2006. 66:290–302. Google Scholar

18.

J Maldonado-Saldivia, B Funke, RK Pandita, T Schuler, BE Morrow, and H Schorle . Expression of Cdcrel-1 (Pnutl1), a gene frequently deleted in velo-cardio-facial syndrome/DiGeorge syndrome. Mech Dev 2000. 96:121–124. Google Scholar

19.

U Grieshammer, C Cebrian, R Ilagan, E Meyers, D Herzlinger, and GR Martin . FGF8 is required for cell survival at distinct stages of nephrogenesis and for regulation of gene expression in nascent nephrons. Development 2005. 132:3847–3857. Google Scholar

20.

N Geijsen, M Horoschak, K Kim, J Gribnau, K Eggan, and GQ Daley . Derivation of embryonic germ cells and male gametes from embryonic stem cells. Nature 2004. 427:148–154. Google Scholar

21.

U Werling and H Schorle . Conditional inactivation of transcription factor AP-2gamma by using the Cre/loxP recombination system. Genesis 2002. 32:127–129. Google Scholar

22.

MD Tallquist and P Soriano . Epiblast-restricted Cre expression in MORE mice: a tool to distinguish embryonic vs. extra-embryonic gene function. Genesis 2000. 26:113–115. Google Scholar

23.

H Lomeli, V Ramos-Mejia, M Gertsenstein, CG Lobe, and A Nagy . Targeted insertion of Cre recombinase into the TNAP gene: excision in primordial germ cells. Genesis 2000. 26:116–117. Google Scholar

24.

W Gu, S Tekur, R Reinbold, JJ Eppig, YC Choi, JZ Zheng, MT Murray, and NB Hecht . Mammalian male and female germ cells express a germ cell-specific Y-Box protein, MSY2. Biol Reprod 1998. 59:1266–1274. Google Scholar

25.

K Yomogida, H Ohtani, H Harigae, E Ito, Y Nishimune, JD Engel, and M Yamamoto . Developmental stage- and spermatogenic cycle-specific expression of transcription factor GATA-1 in mouse Sertoli cells. Development 1994. 120:1759–1766. Google Scholar

26.

Y Wang, DC Newton, TL Miller, AM Teichert, MJ Phillips, MS Davidoff, and PA Marsden . An alternative promoter of the human neuronal nitric oxide synthase gene is expressed specifically in Leydig cells. Am J Pathol 2002. 160:369–380. Google Scholar

27.

AM Elliott, MP de Miguel, VI Rebel, and PJ Donovan . Identifying genes differentially expressed between PGCs and ES cells reveals a role for CREB-binding protein in germ cell survival. Dev Biol 2007. 311:347–358. Google Scholar

28.

S Hayashi and AP McMahon . Efficient recombination in diverse tissues by a tamoxifen-inducible form of Cre: a tool for temporally regulated gene activation/inactivation in the mouse. Dev Biol 2002. 244:305–318. Google Scholar

29.

J Kehler, E Tolkunova, B Koschorz, M Pesce, L Gentile, M Boiani, H Lomeli, A Nagy, KJ McLaughlin, HR Scholer, and A Tomilin . Oct4 is required for primordial germ cell survival. EMBO Rep 2004. 5:1078–1083. Google Scholar

30.

H Schorle, P Meier, M Buchert, R Jaenisch, and PJ Mitchell . Transcription factor AP-2 essential for cranial closure and craniofacial development. Nature 1996. 381:235–238. Google Scholar

31.

Y Mizuno, A Gotoh, S Kamidono, and S Kitazawa . Establishment and characterization of a new human testicular germ cell tumor cell line (TCam-2) [in Japanese]. Nippon Hinyokika Gakkai Zasshi 1993. 84:1211–1218. Google Scholar

32.

D Eckert, K Biermann, D Nettersheim, AJ Gillis, K Steger, HM Jack, AM Muller, LH Looijenga, and H Schorle . Expression of BLIMP1/PRMT5 and concurrent histone H2A/H4 arginine 3 dimethylation in fetal germ cells, CIS/IGCNU and germ cell tumors. BMC Dev Biol 2008. 8:106..  Google Scholar

33.

D Eckert, D Nettersheim, LC Heukamp, S Kitazawa, K Biermann, and H Schorle . TCam-2 but not JKT-1 cells resemble seminoma in cell culture. Cell Tissue Res 2008. 331:529–538. Google Scholar

34.

J de Jong, H Stoop, AJ Gillis, RJ van Gurp, GJ van de Geijn, M Boer, R Hersmus, PT Saunders, RA Anderson, JW Oosterhuis, and LH Looijenga . Differential expression of SOX17 and SOX2 in germ cells and stem cells has biological and clinical implications. J Pathol 2008. 215:21–30. Google Scholar

35.

J de Jong, H Stoop, AJ Gillis, R Hersmus, RJ van Gurp, GJ van de Geijn, E van Drunen, HB Beverloo, DT Schneider, JK Sherlock, J Baeten, S Kitazawa, EJ van Zoelen, K van Roozendaal, JW Oosterhuis, and LH Looijenga . Further characterization of the first seminoma cell line TCam-2. Genes Chromosomes Cancer 2008. 47:185–196. Google Scholar

36.

K Ancelin, UC Lange, P Hajkova, R Schneider, AJ Bannister, T Kouzarides, and MA Surani . Blimp1 associates with Prmt5 and directs histone arginine methylation in mouse germ cells. Nat Cell Biol 2006. 8:623–630. Google Scholar

37.

JB Burch . Regulation of GATA gene expression during vertebrate development. Semin Cell Dev Biol 2005. 16:71–81. Google Scholar

38.

LL Tres, C Rosselot, and AL Kierszenbaum . Primordial germ cells: what does it take to be alive? Mol Reprod Dev 2004. 68:1–4. Google Scholar

39.

S Gajavelli, PM Wood, D Pennica, SR Whittemore, and P Tsoulfas . BMP signaling initiates a neural crest differentiation program in embryonic rat CNS stem cells. Exp Neurol 2004. 188:205–223. Google Scholar

40.

T Luo, YH Lee, JP Saint-Jeannet, and TD Sargent . Induction of neural crest in Xenopus by transcription factor AP2alpha. Proc Natl Acad Sci U S A 2003. 100:532–537. Google Scholar

41.

T Luo, M Matsuo-Takasaki, ML Thomas, DL Weeks, and TD Sargent . Transcription factor AP-2 is an essential and direct regulator of epidermal development in Xenopus. Dev Biol 2002. 245:136–144. Google Scholar

42.

Y Zhang, T Luo, and TD Sargent . Expression of TFAP2beta and TFAP2gamma genes in Xenopus laevis. Gene Expr Patterns 2006. 6:589–595. Google Scholar

43.

KL Loveland and S Schlatt . Stem cell factor and c-kit in the mammalian testis: lessons originating from Mother Nature's gene knockouts. J Endocrinol 1997. 153:337–344. Google Scholar

44.

M Saitou, SC Barton, and MA Surani . A molecular programme for the specification of germ cell fate in mice. Nature 2002. 418:293–300. Google Scholar

45.

Y Yabuta, K Kurimoto, Y Ohinata, Y Seki, and M Saitou . Gene expression dynamics during germline specification in mice identified by quantitative single-cell gene expression profiling. Biol Reprod 2006. 75:705–716. Google Scholar

46.

Y Seki, K Hayashi, K Itoh, M Mizugaki, M Saitou, and Y Matsui . Extensive and orderly reprogramming of genome-wide chromatin modifications associated with specification and early development of germ cells in mice. Dev Biol 2005. 278:440–458. Google Scholar

47.

M Yamaji, Y Seki, K Kurimoto, Y Yabuta, M Yuasa, M Shigeta, K Yamanaka, Y Ohinata, and M Saitou . Critical function of Prdm14 for the establishment of the germ cell lineage in mice. Nat Genet 2008. 40:1016–1022. Google Scholar

Notes

[1] Financial disclosure Supported by the DFG to H.S. (No. 503/6, 503/7, and 503/9).

[2] Equal contributor These authors contributed equally to this work.

Susanne Weber, Dawid Eckert, Daniel Nettersheim, Ad J.M. Gillis, Sabine Schäfer, Peter Kuckenberg, Julia Ehlermann, Uwe Werling, Katharina Biermann, Leendert H.J. Looijenga, and Hubert Schorle "Critical Function of AP-2gamma/TCFAP2C in Mouse Embryonic Germ Cell Maintenance," Biology of Reproduction 82(1), 214-223, (23 September 2009). https://doi.org/10.1095/biolreprod.109.078717
Received: 7 May 2009; Accepted: 1 September 2009; Published: 23 September 2009
KEYWORDS
AP-2γ
developmental biology
gamete biology
gene regulation
HOXB1
PRDM1
primordial germ cells
Back to Top